Chiral Emission into Nanophotonic Resonators | ACS Photonics

Chiral emission, where the handedness of a transition dipole determines the direction ... a photon is emitted, has recently been observed from atoms a...
0 downloads 0 Views 3MB Size
Subscriber access provided by WEBSTER UNIV

Article

Chiral emission into nanophotonic resonators Diego Martin Cano, Harald Haakh, and Nir Rotenberg ACS Photonics, Just Accepted Manuscript • DOI: 10.1021/acsphotonics.8b01555 • Publication Date (Web): 26 Feb 2019 Downloaded from http://pubs.acs.org on February 27, 2019

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 19 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Photonics

Chiral emission into nanophotonic resonators Diego Martin-Cano,∗,† Harald R. Haakh,†,¶ and Nir Rotenberg∗,‡ †Max Planck Institute for the Science of Light, Staudtstr. 2, D-91058 Erlangen, Germany ‡Niels Bohr Institute and Center for Hybrid Quantum Networks (Hy-Q), University of Copenhagen, Blegdamsvej 17, DK-2100 Copenhagen, Denmark ¶Past address E-mail: [email protected]; [email protected] Abstract Chiral emission, where the handedness of a transition dipole determines the direction in which a photon is emitted, has recently been observed from atoms and quantum dots coupled to nanophotonic waveguides. Here, we consider the case of chiral light-matter interactions in resonant nanophotonic structures, deriving closed-form expressions for the fundamental quantum electrodynamic quantities that describe these interactions. We show how parameters such as the position dependent, directional Purcell factors and mode volume can be calculated using computationally efficient two dimensional eigenmode simulations. As an example, we calculate these quantities for a prototypical ring resonator with a geometric footprint of only 4.5 µm2 , observing perfect directionality with a simultaneous Purcell enhancement upwards of 400. The ability to determine these fundamental properties of nanophotonic chiral interfaces is crucial if they are to form elements of quantum circuits and networks.

Keywords Quantum nanophotonics, chiral optics, photonic resonators 1

ACS Paragon Plus Environment

ACS Photonics 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The emergence of chiral light-matter interactions in nanophotonic systems has redefined our understanding of fundamental processes such as the emission or scattering of photons by quantum emitters. 1 In nanophotonic structures the spin and momentum of a local light fields can be locked, 2 meaning that at nanoscopic length scales the direction in which the orientation vector of the light field rotates depends on the direction in which photons travel. As a direct consequence, counter-propagating modes will interact differently with emitters whose transitions are described by elliptically polarized transition dipoles, resulting in highly unidirectional emission or scattering as sketched in Fig. 1. Interestingly, this type of chiral light-matter interaction is a purely electronic phenomenon, and should not be confused with the emission from intrinsically chiral structures such as molecules where, for example, circular dichroism arises due to a combined electric and magnetic response. 3 Electric chiral interactions have recently been observed in a variety of classical 4,5 and quantum 6–8 nanophotonic systems, including in optical nanofibers, nanobeams and photonic crystal waveguides. The existence of these chiral quantum elements has motivated a number of recent proposals that include new routes toward robust generation of entanglement, 9,10 nonreciprocal circuits 7,11 and quantum processing with chiral networks. 12,13 Moreover, our growing understanding of chiral effects has led to novel technological solutions to outstanding photonic and quantum optical challenges, such as single-photon switches 14 and optical isolators, 15 and the demonstration of robust, path-dependent preparation of the spin of a single quantum dot. 16 To operate at the single photon level, researchers have recently turned to cavity-based chiral quantum interfaces that direct photons from a quantum emitter primarily into one of two counter-propagating modes, while enhancing this chiral light-matter interaction. 14,17–19 In a whispering gallery mode resonator (WGM), such as the ring illustrated in Fig. 1, this means that the rate of emission into the left-handed (Γ− ) and right-handed (Γ+ ) degenerate modes differ. This, in turn, means that the total probability of emitting a single photon into

2

ACS Paragon Plus Environment

Page 2 of 19

Page 3 of 19 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Photonics

(ii) Left handed

*-

emission

(i) Right handed

E

J

2

(iii)

z r

emission

*-

*+

Emitter O

*+

z^

R I ^

r^

Figure 1: Chiral emission into a whispering-gallery-mode photonic resonator. The main figure sketches the emission of a right-handed transition dipole located on the outer edge of a ring resonator, where a fast emission rate into the right-handed mode (Γ+ ) and slow emission rate into the left-handed mode (Γ− ) are shown with a thick red line and a thin blue line, respectively. All other losses are characterized by the rate γ. The asymmetric emission occurs because, at the position of the emitter, the polarization of the electric field of the right-handed mode is primarily right-handed (and hence overlaps well with the dipole vector, inset i.). Conversely, the field of the left-handed mode, at the position of the emitter, is primarily left-handed polarized (inset ii.). The size of the circles, in insets i. and ii., indicate the fraction of the electric field that is right-handed polarized and would therefore overlap with the transition dipole moment. For completeness, inset iii. shows the amplitude of the electric field distribution of a ring resonator with radius R = 1.19 µm, 162 nm width and 210 nm height (refractive index n = 3.2 of gallium phosphide, surrounded by air), with maxima observed both inside and outside the waveguide.

each mode (β± ) is modified, since

β± =

Γ± , Γ+ + Γ− + γ

(1)

where γ is the rate at which light couples to all other modes via radiative and nonradiative processes. The viability of a chiral quantum interface is to a large extent determined by the losses, which ideally should be very small: γ  Γ+ + Γ− . In practice, this requirement can be met either by suppressing radiative losses through careful engineering of waveguides 20 or, increasingly, through the enhancement of emission into the desired modes using resonant structures. 21,22 The latter approach shows much promise, particularly with solid-state emitters, as it can compensate for radiation losses and dephasing due to the phononic envi3

ACS Paragon Plus Environment

ACS Photonics 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 4 of 19

ronment, and provide higher photon count rates. 23 Historically, this type of enhancement of the emission of linear dipoles has been characterized by the Purcell factor, F = Γ/Γ0 , where Γ is the decay rate into a single resonant mode and Γ0 is the emission rate into a homogenous material hosting the emitter. 24 Finding an analogous quantum electrodynamical quantity for chiral emission is critical if we are to understand the ultimate limits of chiral quantum effects and devices. Here we address this need, providing closed-form expressions for Γ± and F± for nanophotonic resonators and, in the process, introducing the concept of a chiral mode volume V± . This allows us to calculate the asymmetric emitter-photon coupling efficiencies β± for realistic structures, a crucial ingredient in the description of chiral photonic devices 18,19 (see the Supplementary Information for the link between V± and F± and emitter-resonator coupling constant g± used in the full quantum calculations of these devices).

1 1.1

Formalism Chiral nanophotonic resonators

As is the case for linear waveguides, a resonator-based chiral quantum interface demands the presence of two independent, counter-propagating modes whose complex frequencies are degenerate ω+ = ω− = ω ˜ . 25 In practise, this precludes the use of Fabry-Perot cavities, rather requiring the use of WGM resonators such as bottle, disk or ring microstructures. These support two degenerate quasi-normal modes (QNM) with opposite propagation directions, which are solutions of the inhomogeneous Maxwell Equations. These QNMs, so-called since they can experience both radiation and absorption losses, 26 can be expressed by electric and magnetic field distributions, e.g.

˜ ± (r) = E ¯ ± (r, z) e∓iqϕ . E

4

ACS Paragon Plus Environment

(2)

Page 5 of 19 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Photonics

Here, we have explicitly accounted for the symmetry of the microresonators under consideration by separating the longitudinal (i.e. the azimuthal variable ϕ, see Fig. 1) dependency from the transverse one, noting that the mode must fit an integer number q of waves into the ¯ + (r, z) 2 is illustrated resonator. A prototypical transverse intensity distribution of a QNM E in the inset of Fig. 1, where the confinement of the light to the high-index guiding region ˜+ = E ˜ ∗ ,1 is clearly visible. Time reversal symmetry of Maxwell’s Equations ensures that E − where the asterisk indicates complex conjugation. For the WGM resonators this results in ¯ ± (r, z): 27 the following relations for the components E E¯r,+ = −E¯r,−

E¯z,+ = −E¯z,−

E¯ϕ,+ = E¯ϕ,−

¯ r,+ = H ¯ r,− H

¯ z,+ = H ¯ z,− H

¯ ϕ,+ = −H ¯ ϕ,− H

(3)

The change of sign of the transverse electric field components ensures that the polarization ellipse traced out by the Er and Eϕ components of the counter-propagating QNMs rotate in the opposite directions. 28 With these properties of the QNMs of the WGM resonator, we can now set out to derive expressions for Γ± , V± and, ultimately, F± .

1.2

General expressions for chiral emission rates

We use the first-principle approach introduced by Sauvan et al., 29 which is based on the QNM expansion of Maxwell’s Equations and Fermi’s golden rule. Here the electromagnetic field emitted by an electric dipole in a nanophotonic system is expanded as a sum of QNMs,

E (r, ω) ≈

M X

˜ m (r) , αm (ω) E

(4)

m=1

where αm (ω) are the complex coefficients that describe the amplitude and relative phase that each QNM contributes. A similar expression may be written for the magnetic field.

5

ACS Paragon Plus Environment

ACS Photonics 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 6 of 19

Each QNM fulfills the Lorentz Reciprocity Theorem, Z

n ˜n d3 r E · [ωε (ω) − ω ˜ n ε (˜ ωn )] E o ˜n , − H · [ωµ (ω) − ω ˜ n µ (˜ ωn )] H

˜ n (r0 ) = −ωp · E

(5)

that relates the electric field radiated by a dipole moment p located at position r0 into one QNM at frequency ωn to the overlap between that QNM and the total electromagnetic field at frequency ω. Here, ε = 0 εr and µ = µ0 µr are the permittivity and permeability characterizing the inhomogeneous nanophotonic system. These magnitudes vary slowly as a function of frequency over the typical narrow bandwidths of microresonators and emitters, so that dispersion is neglected. 30 Eq. 4 can be used to reformulate Eq. 5 as a linear system of equations, X

˜ n (r0 ) , Anm (ω) (ω − ω ˜ m ) αm (ω) = −ωp · E

(6)

m

where (ω − ω ˜n) Anm (ω) = (ω − ω ˜m)

Z

i h ˜ ˜ ˜ ˜ d r E m · εE n − H m · µ H n . 3

(7)

Solving these equations yields the coefficients αm (ω) of the QNM expansion of Eq. 4. In obtaining these, we diverge from the work of Sauvan et al., who treated the case of normal emission from linear dipoles 29 into individual non-degenerate modes. In contrast, the asymmetric chiral emission considered here requires the explicit treatment of complex-valued dipole vectors and degenerate counter-propagating modes, meaning that both coefficients, α+ and α− , must be calculated. 31 In this (+, −) basis, A++ = A−− = 0 due to the azimuthal symmetry of the QNM (Eq. 2) and the volume integral in Eq. 7. Consequently, we can write Z A+− (ω) = A−+ (ω) =

h i ˜ + · εE ˜− − H ˜ + · µH ˜− . d3 r E

6

ACS Paragon Plus Environment

(8)

Page 7 of 19 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Photonics

This expression is used to invert A and applied in conjunction with Eq. 6 to obtain

α± = −

ωp · E∓ (r0 ) . (ω − ω ˜ )A+− (ω)

(9)

Fermi’s Golden Rule can relate the total decay rate to contributions of the individual QNMs through " # X 2 ˜ m (r0 ) , Γ = Im α m p∗ · E h ¯ m

(10)

so that the emission rate into each of the counter-propagating modes is h ih i  ˜ ∓ (r0 ) p∗ · E ˜ ± (r0 )  p · E 2 h i Γ± = − ωIm . R  (ω − ω h ¯ ˜ + · εE ˜− − H ˜ + · µH ˜−  ˜ ) d3 r E  

(11)

This compact equation captures the physics of chiral emission since it enables asymmetric directional coupling, i.e. Γ+ 6= Γ− : if, for example, there is a large overlap between p∗ and ˜ + such that Γ+ is maximized, there will be little overlap between p∗ and E ˜ − and thus Γ− E ˜ + and E ˜ −, will be slow. It is interesting to note that while Γ+ (and Γ− ) depends on both E the factor with which each mode modifies the emission rate differs for elliptical dipoles. Equation 11 can be normalized to the emission of an equivalent dipole into bulk media of refractive index n Γ0 = ω 3 |p|2 n/ (3π0 h ¯ c3 ), giving the on-resonance chiral emission enhancement due to the WGM resonator, i.e. the directional Purcell factors 3 Γ± = 2 F± = Γ0 4π



λ0 n

3

 Re

Q V±

 .

(12)

Here, ω0 = Re ω ˜ and Q = Re ω ˜ / (2Im ω ˜ ) are the central frequency and quality factor of the resonance, and we have also defined the position-dependent chiral mode volumes to be h i ˜ + · εE ˜− − H ˜ + · µH ˜− d3 r E h ih i, V± (r0 ) = ∗ ˜ ˜ 2ε u · E∓ (r0 ) u · E± (r0 ) R

7

ACS Paragon Plus Environment

(13)

ACS Photonics 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

where u is the unit vector of the dipole (p = pu). That these mode volumes depend on the position of the emitter might, initially appear odd. Yet, very similarly, the Purcell factor acquires positional dependence through the local density of optical states if a Green’s tensor approach is chosen instead. 32 This position-dependent normalization factor arises directly from our derivation and provides a clear and consistent way to define chiral volumes, particularly for lossy modes and in the presence of absorption, where the usual formulation fails. 26,33 Finally, and for completeness, we note that although quantities such as decay rates or the Purcell factor are fundamentally quantum in nature they can in practise be extracted from classical calculations. Classically, the power radiated by a dipole in a given photonic environment is related to the imaginary component of the Greens tensor. Within the weak coupling regime, this quantity is proportional to the quantum decay rate enhancement. Moreover, within the macroscopic quantum electrodynamics formalism the cavity coupling parameters can even be estimated in the strong coupling regime, if both the Greens tensor and the mode volume are known 34 (see the Supporting information for more details).

2

Resonant chiral emission

We now apply our theory to a realistic nanophotonic system that can support and enhance chiral emission, namely the ring resonator sketched out in Fig. 1. Due to the azimuthal ˜ ± can be determined from two-dimensional symmetry of WGM resonators, the QNMs E eigenmode calculations, and so the problem reduces to solving for the different αm ’s. As an example, we consider a ring with a radius of only 1.19 µm, of rectangular cross-section of 210 nm height and 162 nm width, and made out of a gallium phosphide (n = 3.2) surrounded by air. For this geometry, the azimuthal mode of order g = 21 is found at a frequency of 393.58 THz, with a moderate Q ≈ 20, 000, well within the range of what can be realistically fabricated. 35

8

ACS Paragon Plus Environment

Page 8 of 19

Page 9 of 19

400

15.5

100

155

10

1550

1

Mode volume (V /λ 3)

z 100 nm

r

15500

waveguide

400

waveguide

300 200 100 1000

1100

1200

r (nm)

1300

1400

1000

1100

1200

r (nm)

1300

1400

0

Purcell factor (F )

3.9

Γ+

Γ-

0.1

Purcell factor (F )

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Photonics

Figure 2: Spatial dependence of the chiral mode volumes and Purcell factors for emission into the two counter-propagating modes of a ring-resonator, calculated for a right-hand circular √ dipole u+ = 1/ 2 (ˆ r + iϕ) ˆ on a cross-section of the waveguide (dashed square, inner edge to ˜ − , top left panel) occurs mainly on the inner the left). Emission into the left-handed mode (E ˜ + , bottom right panel) side of the waveguide, while emission into the right-handed mode (E occurs mainly on the outer side of the waveguide. For clarity, line cuts of the corresponding Purcell factors along the symmetry axis (dashed line in top panels) are shown in the bottom panels, on a linear scale. An opposite-handed dipole, u− , will emit in the opposite fashion (i.e. a swapping of the panels). The line along which β± is calculated for Fig. 3 below is shown in solid in the top panels, and denoted by the vertical dashed lines in the bottom panels.

We calculate the vectorial components of this mode using a finite-element eigenmode solver (Comsol Multiphysics), showing the resultant intensity distribution in the inset of Fig. 1. Equations 12 and 13, in conjunction with the relations between the counter-propagating fields (Eq. 3) allow us to very simply map out the spatial dependence of the chiral mode volumes and Purcell factors, which we show in Fig. 2 for a right-handed circular dipole √ u+ = (ˆ r + iϕ) ˆ / 2. The difference between V+ and V− , and hence between F+ and F− , is ˜ + mode when it is placed near the outer (right) striking: The emitter radiates into the E ˜ − mode occurs primarily when it is found at the side of the bar, while emission into the E 9

ACS Paragon Plus Environment

ACS Photonics 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

inner (left) side. Note, too, that V+ is not simply a mirror image of V− , as a slight asymmetry is introduced between these two distributions by the bending of the ring, which results in a higher mode intensity and hence emission enhancement at the outer side of the ring. √ Changing the dipole polarization rotation to u− = (ˆ r − iϕ) ˆ / 2 simply reverses the emission direction (i.e. exchanges the panels of Fig. 2). For both modes, we observe that emission peaks first inside the waveguide and then again, with greater amplitude outside, near its interface. For this structure we calculate a maximal F+ ≈ 80 inside and F+ ≈ 400 outside the bar, meaning that this specific geometry is highly advantageous for quantum emitters that are embedded in the low-index material surrounding the ring, such as organic molecules in a solid-state matrix or trapped atoms. Tuning the ring geometry, either by increasing the bar cross-section or the radius of the ring, pushes the mode into the high-index material and hence enhances the emission inside the waveguide, as would be favourable for emitters such as defect centers in diamond or quantum dots. The directional nature of the emission of the circular dipole is clear from Fig. 2 since regions of high F+ do not overlap with regions of high F− . We quantify this directionality in Fig. 3(a), where we plot the normalised difference of the rate of emission between Γ+ and Γ− for a right-handed circular dipole u+ . Here, we observe that perfect directional emission is possible for this small ring resonator with regions of +1 directionality located inside the resonator on its outer side, while regions of where the directionality is -1 can be found both inside and outside the ring in its inner side. As before, this asymmetry is due to the bending of the waveguide, which pushes the fields to the outer side of the ring. Our theory can also be used to quantify the directional coupling efficiencies β± , which determine how many photons are emitted into each mode of the ring. Importantly, as we discuss below, our approach allows for the calculation of β± without resorting to powerflux monitors, which greatly complicate the calculations in the presence of large radiation losses or absorption. To begin with, however, when γ is negligible, as is the case for our

10

ACS Paragon Plus Environment

Page 10 of 19

Page 11 of 19

(a ) -1 .0

-0 .5

D ir e c tio n a lity 0 .0

0 .5

1 .0

F +

- F

F +

+ F

-

1 0 0 n m

(b )

1 .0 0

+

,-

0 .7 5



1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Photonics

0 .5 0

0 .2 5

0 .0 0 1 1 0 0

1 1 5 0

1 2 0 0

1 2 5 0

1 3 0 0

1 3 5 0

D ip o le p o s itio n ( n m )

Figure 3: The emission of a right-handed dipole u+ into the ring resonator characterized in Fig. 2. (a) Spatial map of the directionality of emission, which is seen to range from +1 (all emission into the right-handed mode; outer side of the resonator) to -1 (all emission into the left-handed mode; inner side of the resonator). (b) The directional coupling efficiencies calculated for dipole positions along the axis of symmetry of the waveguide. The symbols are the β± extracted from full-vectorial three-dimensional calculations, while the solid curves are found using Eqs. 1, 12 and 13, and a single, two-dimensional eigenmode calculation. The shaded region denotes the waveguide.

ring resonator, we can immediately use Eq. 1. We plot β± along the symmetry axis of the waveguide (solid lines in Fig. 2) in Fig. 3(b), observing excellent agreement between our theory (curves) using our single, eigenmode calculation with Eq. 12 and full-vectorial threedimensional calculations (symbols, see below for explanation). For this ring, both β+ and β− assume peak values of ≈ 0.99 inside the waveguide and, in the region that we consider, a maximum value of β+ ≈ 0.96 occurs just to the outside. Conversely, we also observe a 11

ACS Paragon Plus Environment

ACS Photonics 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

crossing point near r ≈ 1200 nm, where β+ = β− and hence the emission is bidirectional. If other significant decay channels are available to the emitter, for example through radiation into free-space, then γ must also be quantified if β± are to be determined. 36 In practise this can be done in two ways, using the three-dimensional simulations: 22 (i) Either, one calculates βtot , by fitting from the leaky mode envelope, which then allows for the calculation of β± = βtot F± / (F+ + F− ), where F± is found using Eq. 12. (ii) Alternatively, the total emission enhancement η, which is defined to be the ratio of the power radiated by the dipole in the resonator to that of the dipole in bulk, and which in light of Eq. 1, is related to the directional coupling efficiencies through β± = F± /η. Approach (ii), in particular, does not require any fitting and provides a clear and unambiguous way to calculate the directional emission efficiencies numerically. In summary, we note that even in the presence of large losses, F± can be extracted from two-dimensional eigenmode calculations via Eq. 12; threedimensional simulations are required only for β± (See Supporting Information for examples of ring resonators with large losses due to either a poor mode confinement or, in the case of plasmonics, absorption).

3

Conclusions and outlook

The ability to accurately and efficiently quantify the interactions of quantum emitters with nanophotonic structures is key if the promise of theoretical proposals and proof-of-concept experiments is to be fulfilled. For chiral quantum interactions, we have provided a framework within which quantum electrodynamic parameters such as directional decay rates and coupling strengths, can be calculated for realistic and compatible photonic resonators. As an example, we have calculated these parameters for a dielectric ring resonator with computationally favorable 2D models, and observed excellent agreement between our theory and full-numerical 3D simulations. We have shown that a careful positioning of an emitter in this simple resonator allows for any degree of directionality, all with a high Purcell enhancement

12

ACS Paragon Plus Environment

Page 12 of 19

Page 13 of 19 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Photonics

and near-unity coupling efficiency. This demonstrates the potential of our approach to design and engineer realistic and efficient quantum chiral interfaces, representing an important step towards the realization of quantum photonic processing. 1,12,13 Our theory also describes classical chiral light-matter interactions 4,5 and could therefore open new routes towards creating and designing chiral photonic elements, and enhancing the spectroscopy and sensing of chiral molecules. 37

Acknowledgement The author thank Pierre T¨ urschmann and Peter Lodahl for stimulating discussions.

Funding Sources This work was supported by the Max Planck Society.

Supporting Information Available Introduction of the Jaynes-Cumming model for linear dipoles in cavity QED and its relation to the mode volume for open resonators. Link of F± and V± to chiral cavity QED coupling parameters in open resonators. Examples of lossy chiral WGM resonators: poor-confined dielectric and plasmonic rings.

References (1) Lodahl, P.; Mahmoodian, S.; Stobbe, S.; Rauschenbeutel, A.; Schneeweiss, P.; Volz, J.; Pitchler, H.; Zoller, P. Chiral quantum optics. Nature 2017, 541, 473–480. (2) Bliokh, K. Y.; Bekshaev, A. Y.; Nori, F. Extraordinary momentum and spin in evanescent waves. Nature Commun. 2014, 5, 3300. 13

ACS Paragon Plus Environment

ACS Photonics 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(3) Yoo, S.; Park, Q.-H. Chiral light-matter interaction in optical resonators. Phys. Rev. Lett. 2015, 114, 203003. (4) Petersen, J.; Volz, J.; Rauschenbeutel, A. Chiral nanophotonic waveguide interface based on spin-orbit interaction of light. Science 2014, 346, 67–71. (5) le Feber, B.; Rotenberg, N.; Kuipers, L. Nanophotonic control of circular dipole emission. Nat. Commun. 2015, 6, 6695. (6) Mitsch, R.; Sayrin, C.; Albrecht, B.; Schneeweiss, P.; Rauschenbeutel, A. Quantum state-controlled directional spontaneous emission of photons into a nanophotonic waveguide. Nat. Commun. 2014, 5, 5713. (7) S¨ollner, I.; Mahmoodian, S.; Hansen, S. L.; Midolo, L.; Javadi, A.; Kirˇsansk˙e, G.; Pregnolato, T.; El-Ella, H.; Lee, E. H.; Song, J. D.; Stobbe, S.; Lodahl, P. Deterministic photonemitter coupling in chiral photonic circuits. Nat. Nanotech. 2015, 10, 775–778. (8) Coles, R. J.; Price, D. M.; Dixon, J. E.; Royall, B.; Clarke, E.; Kok, P.; Skolnick, M. S.; Fox, A. M.; Makhonin, M. N. Chirality of nanophotonic waveguide with embedded quantum emitter for unidirectional spin transfer. Nat. Commun. 2016, 7, 11183. (9) Ramos, T.; Pichler, H.; Daley, A. J.; Zoller, P. Quantum spin dimers from chiral dissipation in cold-atom chains. Phys. Rev. Lett. 2014, 113, 237203. (10) Gonzalez-Ballestero, C.; Gonzalez-Tudela, A.; Garcia-Vidal, F. J.; Moreno, E. Chiral route to spontaneous entanglement generation. Phys. Rev. B 2015, 92, 155304. (11) Gonzalez-Ballestero, C.; Moreno, E.; Garcia-Vidal, F. J.; Gonzalez-Tudela, A. Nonreciprocal few-photon routing schemes based on chiral waveguide-emitter couplings. Phys. Rev. A 2016, 94, 063817. (12) Pichler, H.; Ramos, T.; Daley, A. J.; Zoller, P. Quantum optics of chiral spin networks. Phys. Rev. A 2015, 91, 042116. 14

ACS Paragon Plus Environment

Page 14 of 19

Page 15 of 19 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Photonics

(13) Mahmoodian, S.; Lodahl, P.; Sørensen, A. S. Quantum networks with chiral-lightmatter interaction in waveguides. Phys. Rev. Lett. 2016, 117, 240501. (14) Shomroni, I.; Rosenblum, S.; Lovsky, Y.; Bechler, O.; Guendelman, G.; Dayan, B. Alloptical routing of single photons by a one-atom switch controlled by a single photon. Science 2014, 345, 903. (15) Scheucher, M.; Hilico, A.; Will, E.; Volz, J.; Rauschenbeutel, A. Quantum optical circulator controlled by a single chirally coupled atom. Science 2016, 354, 1577–1580. (16) Coles, R. J.; Price, D. M.; Royall, B.; Clarke, E.; Skolnick, M. S.; Fox, A. M.; Makhonin, M. N. Path-dependent initialization of a single quantum dot exciton spin in a nanophotonic waveguide. Phy. Rev. B 2017, 95, 121401(R). (17) Junge, C.; OShea, D.; Volz, J.; Rauschenbeutel, A. Strong coupling between single atoms and nontransversal photons. Phys. Rev. Lett. 2013, 110, 213604. (18) Lenferink, A. J.; Wei, G.; Stern, N. P. Coherent optical non-reciprocity in axisymmetric resonators. Opt. Express 2014, 22, 16099. (19) Xia, K.; Lu, G.; Lin, G.; Cheng, Y.; Niu, Y.; Gong, S.; Twamley, J. Reversible nonmagnetic single-photon isolation using unbalanced quantum coupling. Phys. Rev. A 2014, 90, 043802. (20) Arcari, M.; S¨ollner, I.; Javadi, A.; Lindskov Hansen, S.; Mahmoodian, S.; Liu, J.; Thyrrestrup, H.; Lee, E. H.; Song, J. D.; Stobbe, S.; Lodahl, P. Near-unity coupling efficiency of a quantum emitter to a photonic crystal waveguide. Phys. Rev. Lett. 2014, 113, 093603. (21) Schneeweiss, P.; Zeiger, S.; Hoinkes, T.; Rauschenbeutel, A.; Volz, J. Fiber ring resonator with a nanofiber section for chiral cavity quantum electrodynamics and multimode strong coupling. Opt. Lett. 2017, 42, 85–88. 15

ACS Paragon Plus Environment

ACS Photonics 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(22) Rotenberg, N.; T¨ urschmann, P.; Haakh, H. R.; Martin-Cano, D.; G¨otzinger, S.; Sandoghdar, V. Small slot waveguide rings for on-chip quantum optical circuits. Opt. Express. 2017, 5, 5397–5414. (23) Wang, D.; Kelkar, H.; Martin-Cano, D.; Utikal, T.; G¨otzinger, S.; Sandoghdar, V. Coherent coupling of a single molecule to a scanning Fabry-Perot microcavity. Phys. Rev. X 2017, 7, 021014. (24) Purcell, E. M. Spontaneous emission probabilities at radio frequencies. Phys. Rev. 1946, 69, 681. (25) Here, the real part of ω ˜ is the resonance frequency while the imaginary component quantifies the damping of the resonator. (26) Lalanne, P.; Yan, W.; Vynck, K.; Sauvan, C.; Hugonin, J.-P. Light interaction with photonic and plasmonic resonances. Laser Photon. Rev. 2018, 12, 1700113. (27) Collin, R. E. Field theory of guided waves, 2nd ed.; Wiely-IEEE Press, New York, 1990. (28) Note that the same is true for the ellipse of the Ez and Eϕ components, and for the magnetic components, where sign of the longitudinal component flips. (29) Sauvan, C.; Hugonin, J. P.; Maksymov, I. S.; Lalanne, P. Theory of the spontaneous optical emission of nanosize photonic and plasmon resonators. Phys. Rev. Lett. 2013, 110, 237401. (30) Note that a generalization to dispersive resonators is nevertheless possible 29 . (31) Assuming that the modes do not spectrally overlap, we need only consider the two counterpropagating modes so that the sum runs over {+, −} and we omit the frequency index for brevity; the calculation can be repeated for each pair of degenerate QNMs. (32) Novotny, L.; Hecht, B. Principles of Nano-Optics, 2nd ed.; Cambridge University Press, New York, 2012. 16

ACS Paragon Plus Environment

Page 16 of 19

Page 17 of 19 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Photonics

(33) Kristensen, P. T.; Ge, R.-C.; Hughes, S. Normalization of quasinormal modes in leaky optical cavities and plasmonic resonators. Phy. Rev. A 2015, 92, 053810. (34) Vogel, W.; Welsch, D.-G. Quantum Optics, 3rd ed.; Wiley-VCH, Weinheim, 2006. (35) Bruch, A. W.; Xiong, C.; Leung, B.; Poot, M.; Han, J.; Tang, H. X. Broadband nanophotonic waveguides and resonators based on epitaxial GaN thin films. Appl. Phys. Lett. 2015, 107, 141113. (36) In the case absorption losses, the non-radiative component of each emission rate, Γnrad = Γ± − Γ±,rad can be simply calculated 29 from the eigenmode and the absorption volume integral Γnrad

2 = h ¯

Z

2 ˜ ˜ d rIm () α+ E+ (r) + α− E− (r) 3

(14)

. (37) Patterson, D.; Schnell, M.; Doyle, J. M. Enantiomer-specific detection of chiral molecules via microwave spectroscopy. Nature 2013, 497, 475.

17

ACS Paragon Plus Environment

ACS Photonics 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Graphical TOC Entry

F+∝ Q V+

F- ∝ Q V-

18

ACS Paragon Plus Environment

Page 18 of 19

Page 19 of 19 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37

ACS Photonics

F+∝ Q V+

F- ∝ Q VACS Paragon Plus Environment